Skip to main content

Full text of "DTIC ADA015899: Solid Propellant Kinetics. V. Fuel-Oxidizer Reaction Rates from Heterogeneous Opposed Flow Diffusion Flame"

See other formats


AD-A015  899 


SOLID  PROPELLANT  KINETICS.  V.  FUEL-OXIDIZER 
REACTION  RATES  FROM  HETEROGENEOUS  OPPOSED  FLOW 
DIFFUSION  FLAME 

C.  M.  Ablow,  et  al 

Stanford  Research  Institute 


Prepared  for: 

Office  of  Naval  Research 

Defense  Contract  Administration  Services  Region 


December  1974 


DISTRIBUTED  BY: 


National  Technical  Information  Service 
U.  S.  DEPARTMENT  OF  COMMERCE 


ADA015899 


296108 


December 


1974 


SOLID  PROPELLANT  KINETICS 


V.  FUEL-OXIDIZER  REACTION  RATES  FROM 
HETEROGENEOUS  OPPOSED  FLOW  DIFFUSION  FLAME*' t 


C.  M.  Ablow  and  H.  Wise 
Stanford  Research  Institute 
Menlo  Park,  California  94025 


Reproduction  in  whole  or  in  part  is  permitted  for 
United  States  Government. 


D D C 

rar?r-D;?iio.  nra 

‘i  i 

^ OCT  16  1975 

iliiLTSEtniE 

Cf  D 

any  purpose  of  the 


t 

This  work  was  sponsored  by  the  Office  of  Naval  Research,  Power  Branch, 
Washington,  D.C. , under  Contract  N00014-70-C-0155. 


Approved  for  public  release;  distribution  unlimited. 


Reprodu't.J  by 

NATIONAL  TECHNICAL 
INFORMATION  SERVICE 


US  Department  of  Commerce 
Springfield,  VA.  22151 


$ 


% 


t 


3 


i 

i 


? 

Jj 

ri 


1 


Unclassified 


SECURITY  CLASSIFICATION  OF  THIS  PAGE  (Whan  Data  Entaradl 


REPORT  DOCUMENTATION  PAGE 


1.  REPORT  NUMBER 


READ  INSTRUCTIONS 
BEFORE  COMPLETING  FORM 


2.  GOVT  ACCESSION  NO.  3.  RECIPIENTS  CATALOG  NUMBER 


4.  t1TLE  (and  Subti(la) 

SOLID  PROPELLANT  KINETICS  V.  FUEL-OXIDIZER 
REACTION  RATES  FROM  HETEROGENEOUS  OPPOSED  FLOW 
DIFFUSION  FLAME 


7.  AUTNOR(j) 

C.  M.  Ablow  and  H.  Wise 


9.  PERFORMING  ORGANIZATION  NAME  AND  ADDRESS 

Stanford  Research  Institute 
Menlo  Park,  California  94025 


11.  CONTROLLING  OFFICE  NAME  ANO  ADDRESS 

Office  of  Naval  Research,  Power  Branch 
Department  of  the  Navy,  Washington,  D.C. 


14.  MONITORING  AGENCY  NAME  & ADDRESS  (if  diff.  from  Controlling  Offica) 

DCASR  - San  Francisco 
866  Malcolm  Road 
Burlingame,  CA  94010 


5.  TYPE  OF  REPORT  & PERIOD  COVERED 

Interim  Report 


6.  PERFORMING  ORG.  REPORT  NUMBER 

PYU-8378 


8.  CONTRACT  OR  GRANT  NUMBERS) 


N00014-70-C-0155 


10.  PROGRAM  ELEMENT.  PROJECT.  TASK 
AREA  & WORK  UNIT  NUMBERS 

NR-092-507 


T2.  REPORT  DATE 

December  1974 


15.  SECURITY  CLASS,  (of  thi*  report) 

Unclassified 


15a.  DECLASSIFICATION /DOWNGRADING 
SCHEDU'.E 


16.  DISTRIBUTION  STATEMENT  (of  thi*  report) 

Distribution  of  this  document 


is  unlimited. 


17.  DISTRIBUTION  STATEMENT  (of  th*  abstract  antartd  in  Block  20,  if  diffarant 


D D C 


19-  KEY  WORDS  (Continue  on  reverse  side  if  necessary  and  identify  by  block  number) 

Diffusion  Flame 
Opposed  Flow 
Propellant  Kinetics 

Ammonium  Perchlorate 
Flame  Model 


20.  ABSTRACT  (Commut  on  reverse  side  if  necessary  and  identify  by  block  number) 

A theoretical  model  is  presented  relating  the  gas  dynamics  and  chemical  kinetics 
of  the  opposed  flow  diffusion  flame  formed  in  the  stagnation  region  between  two 
opposing  streams  of  gaseous  reactants,  one  originating  from  the  surface  of  a sub- 
liming solid,  such  as  ammonium  perchlorate.  At  low  gas  flows  the  regtession  rate 
of  the  solid  is  controlled  by  the  physical  properties  of  the  system,  including  the 
net  heat  of  gasification,  the  heat  of  combustion,  and  the  transport  parameters. 

At  high  gas  flows  a limiting  solid  regression  rate  is  attained  due  to  reaction-rate 


>,STn1473 

ON  OF  1 NOV  65  IS  OBSOLETE 


Unclassif ied 

SECURITY  CLASSIFICATION  OF  THIS  PAGE  (When  D*t»  Entered) 


I 

1 


Unclassified 


FCURITY  CLASSIFICATION  OF  THIS  PAGE  (Whin  Oat*  Enter td) 


19.  KEY  WORDS  (Continued) 


20  ABSTRACT  (Contmuad) 

limitations  that  cause  incomplete  combustion  of  the  reactants.  The  theoretical 
model  developed  for  the  heterogeneous  opposed  flow  diffusion  flame  allows  inter- 
pretation of  the  limit  in  solid  regression  rates  in  terms  of  global  reaction 
kinetics.  Calculations  have  been  carried  out  for  a range  of  parameters,  including 
net  heats  of  gasification  and  activation  energies.  For  the  AP-propylene  system 
the  experimental  data  can  be  fitted  to  a second-order  gas-phase  reaction  rate  with 

an  activation  energy  of  37  ± 1 kcal/mole  and  a preexponential  coefficient  of 
t”?  -i  -l 

10^  cc*mol  ^sec 


DD,  ”“1473'8A™ 

COITION  OF  1 NOV  65  IS  OBSOLETE 


Unclassified 

SECURITY  CLASSIFICATION  OF  THIS  AGE  (Winn  0*1*  E- 


ABSTRACT 


A theoretical  model  is  presented  relating  the  gas  dynamics  and 
chemical  kinetics  of  the  opposed  flow  diffusion  flame  formed  in  the 
stagnation  region  between  two  opposing  streams  of  gaseous  reactants, 
one  originating  from  the  surface  of  a subliming  solid,  such  as  ammonium 
perchlorate.  At  low  gas  flows  the  regression  rate  of  the  solid  is 
controlled  by  the  physical  properties  of  the  system,  including  the  net 
heat  of  gasification,  the  heat  of  combustion,  and  the  transport  param- 
eters. At  high  gas  flows  s limiting  solid  regression  rate  is  attained 
due  to  reaction-rate  limitations  that  cause  incomplete  combustion  of 
the  reactants.  The  theoretical  model  developed  for  the  heterogeneous 
opposed  flow  diffusion  flame  allows  interpretation  of  the  limit  in 
solid  regression  rates  in  terms  of  global  reaction  kinetics.  Calcula- 
tions have  been  carried  out  for  a range  of  parameters,  including  net 
heats  of  gasification  and  activation  energies.  For  the  AP-propylene 
system  the  experimental  data  can  be  fitted  to  a second-order  gas-phase 

reaction  rate  with  an  activation  energy  of  37  ± 1 kcal/mole  and  a 

13  -1  -1 

preexponential  coefficient  of  10  cc*mol  sec 


I 


Introduction 


An  understanding  of  the  solid  propellant  combustion  process  is 
essential  for  interpretation  of  motor  performance  and  the  prediction  of 
propellant  characteristics  under  various  physical  conditions.  To 
elucidate  the  burning  mechanism  of  AP-based  composite  solid  propellant  > 
a number  of  analyses  have  been  carried  out  that  differ  mainly  in  the 
degree  of  complexity  of  the  theoretical  models  employed.  Common  to 
all  thece  models  is  the  mechanism  of  (1)  conductive  heat  transport  to 
the  propellant  surface  from  the  adjoining  high-temperature,  gaseous 
reaction  zone,  (2)  gasification  of  the  condensed-phase  component  (fuel 
and  oxidizer),  (3)  diffuslonal  mixing  of  the  gaseous  constituents,  and 
(4)  exothermic  chemical  reaction  in  the  gas  phase.  Near  the  solid/gas 
interface  an  AP  (ammonium  perchlorate)  decomposition  flame  may  form  part 
of  the  reaction  zone  involving  fuel  and  oxidizer  species  emanating  from 
the  solid  propellant  in  their  original  chemical  composition  or  modified 
by  intervening  chemical  process.  In  addition,  subsurface  reactions  and 
phase  changes  of  the  solid  at  the  surface  may  make  contributions  to  the 
regression  rate  of  the  solid  propellant. 

The  complex  nature  of  the  soli  propellant  reacting  system  and  the 
high  temperatures  involved  make  the  task  of  elucidating  the  numerous  kin- 
etic steps  most  formidable.  However  as  a first  step  in  the  study  of  the 
reaction  mechanism  it  may  be  adequate  to  have  available  information  on 
the  "global"  reaction  kinetics  that  offer  a measure  of  the  heat  release 
rate  due  to  exothermic  chemical  reaction  in  the  gas  phase  as  a function  of 
temperature,  pressure,  and  gas  composition.  The  availability  of  such  data 
should  permit  semiquantitative  interpretation  of  the  reaction  rates  in 
the  flame  zone,  and  provide  a better  measure  of  the  interplay  between 
diffusional  mixing  and  chemical  reaction  under  various  operational  conditions. 

* 1 
References  are  listed  after  the  appendix. 


The  opposed-jet  experimental  technique*"  has  been  used  to  obtain 

kinetics  parameters  In  the  homogeneous  case  where  jets  of  gaseous  fuel 

3 

and  oxidizer  support  a diffusion-controlled  flame.  As  the  velocity  of 
the  jets  is  increased,  a smaller  fraction  of  the  incoming  reactants  are 
consumed  due  to  kinetic  limitations.  The  unburnc  reactants  act  as  diluents 
carrying  away  heat  from  the  flame  so  that  at  a high  enough  velocity  the 
flame  is  extinguished. 

The  same  technique  has  been  applied  to  the  heterogeneous  case  where 

•1 

a jet  of  gaseous  fuel  is  directed  against  a subliming  solid  oxidizer. 

As  the  velocity  of  the  jet  is  increased,  the  solid  regression  rate  has 
been  observed  to  increase  to  a certain  limit,  at  which  a further  increase 
in  jet  velocity  leaves  the  regression  rite  unchanged.  This  limit  has 
been  identified  to  correspond  to  the  maximum  fuel  consumption  rate 
for  the  system.  At  greater  flow  rates,  fuel  slips  through  the  flame 
zone  to  act  as  diluent  and  heat  sink  near  the  subliming  surface,  thus 
limiting  the  surface  temperature  and  regression  rate.  A theoretical  model 
is  developed  that  relates  stagnation  flow  and  regression  rates  to  global 
kinetic  parameters  and  the  heat  requirement  at  the  subliming  surface. 

In  the  following  sections  we  have  prepared  first  a descriptive 
presentation  of  the  model,  tne  approximations  made,  and  the  results 
obtained.  This  part  is  followed  by  Appendix  A containing  the  details  of 
the  mathematical  analysis  and  Appendix  B giving  the  application  10  the 
AP-propyiene  system. 


2 


II  Theoretical  Model 


The  objective  of  the  theoretical  analysis  is  to  relate  the  rates  of 
chemical  reaction  occurring  in  the  gas-phase  diffusion  flame  to  the 
combustion  characteristics  cf  the  heterogeneous  opposed-flow  diffusion 
flame  (HOFD)  set  up  between  a subliming  solid  oxidizer  and  a gaseous 
jet  of  fuel  with  inert  diluent.  Alternatively,  the  analysis  allows 
deduction  of  reaction-rate  parameters  of  the  gas-phase  deflagration 
from  experimental  measurements  of  the  HOFD. 

The  cylindrically  symmetric,  steady  flow  field  is  sketched  in 
Figure  1.  It  has  been  assumed  that  the  ratios  of  jet  nozzle  radius  and 
distance  from  the  solid  to  solid  stick  radius  are  so  large  as  to  be 
effectively  infinite.  Although  there  are  two  coordinates,  radial  distance 
r from  the  axis  of  symmetry  and  axial  distance  z from  the  surface  of  the 
solid,  the  flow  along  the  axis  is  decoupled  from  the  remainder  of  the 
flow  if  one  assumes  that  the  radial  diffusion  of  mass  and  heat  is  negli- 
gible in  comparison  with  axial  diffusion  and  convection.  The  distributions 
of  temperature  and  species  along  the  axis  are  then  determined  by  one- 
dimensional equations  in  z only. 

The  multi-step  kinetics  of  the  gas-phase  deflagration  process  are 
approximated  by  a global  reaction-rate  expression.  Such  an  approximation 
appears  to  be  an  adequate  quantitative  description  of  the  kinetics  in 
terms  of  local  heat-release  rates.  Obviously  it  is  less  satisfactory  in 
terms  of  an  analysis  of  spatial  distribution  of  chemical  intermediates 
since  no  detailed  account  is  taken  of  the  mechanism  of  the  reaction  and 
the  intermediate  species  produced. 

In  Appendix  A the  conservation  equations  for  mass  and  energy  are 
formulated  ir.  terms  of  the  Shvab-Zeldovitch  model  with  constant,  temperature- 
averaged  parameters  for  the  physical  and  transport  parameters,  and  a Lewis 


3 


number  of  unity.  Such  a model  is  based  on  a single  overall  reaction 

involving  two  gaseous  reactants.  The  coefficients  in  the  equations 

depend  on  the  axial  component  of  the  mass-flux  distribution,  which  in 

turn  is  determined  by  the  momentum  balance  equations.  However  in  the 

present  system  the  momentum  variations  are  expected  to  be  small  so  that 

almost  any  flow  satisfying  the  mass  conservation  equation  should  be  an 

5 

adequate  approximation.  Spalding  and  others  have  assumed,  for  the  case 
of  opposed  gaseous  jets,  a linear  variation  of  the  radial  velocity  com- 
ponent with  distance  from  the  axis  and  a similar  variation  of  the  axial 
component  irom  the  plane  through  the  stagnation  point  and  perpendicular 

to  the  axis.  This  flow  can  be  strictly  correct  only  near  the  stagnation 

7 

point.  Seif-similar  boundary  layer  flow  has  also  been  employed.  This 
flow  has  the  advantage  of  providing  a solution  of  the  complete  set  of 
flow  equations.  However,  it  seems  inapplicable  to  the  present  case  where 
a large  regression  rate  disrupts  the  boundary  layer. 

In  the  present  model  the  axial  component  of  the  mass  flux  is  described 
by  a series  approximation  with  a sufficient  number  of  terms  to  satisfy 
the  conditions  at  the  solid  surface  and  at  the  inlet  for  the  gaseous 
reactant.  The  flow  has  a stagnation  point  and  exhibits  the  correct 
behavior  in  its  neighborhood.  The  radial  component  of  the  flow  is  deter- 
mined by  the  axial  mass  flux  gradient  and  has  in  the  model  the  constant 
sign  that  implies  a realistic  radial  outflow  everywhere.  To  provide  an 
analytical  model  for  rapid  computation  and  evaluation  of  different 
parameters  we  approximate  the  temperature  distribution  by  a polynomial 

in  the  distance  variable.  An  advantage  of  the  analytic  approximation 

8 

method  over  asymptotic  methods  is  that  no  parameter  need  take  extreme 
values  for  the  approximation  to  be  useful. 

In  +he  analysis  of  HOFD  information  is  needed  on  the  conditions  at 

the  gas/solid  interface.  For  the  present  model,  we  have  employed  exper- 
10 

imental  results  relating  surface  temperature  and  regression  rate. 


4 


Ill  Analytical  Results 


The  most  striking  feature  of  the  experimental  results  for  the  AP/fuel 
4 

system  is  the  observed  change  from  a strong  variation  of  solid  regression 
rate  at  values  of  the  fuel  mass  flux  to  nearly  constant  regression 
rate  at  higher  fluxes.  The  theoretical  model  interprets  this  behavior 
to  be  due  to  a change  from  complete  fuel  consumption  in  one  case  to 
unburnt  fuel  slippage  through  the  flame  to  the  solid  surface  in  the  other. 
In  the  latter  case  the  unburnt  fuel  acts  as  diluent  and  heat  sink. 


At  very  low  fuel  flow  rates  the  thin-flame  approximation  applies 

(Appendix  A. 9).  In  this  regime  the  combustion  process  is  controlled  by 

the  mass  transfer  rate  of  reactants  to  the  flame  and  is  a sensitive 

function  of  the  heat  L of  gasification  of  the  solid  reactant.  The 

value  of  L is  given  by  the  difference  between  the  heat  of  sublimation 

and  the  exothermic  heat  of  reaction  at  the  solid  surface.  A series  of 

calculations  have  been  carried  out  over  a range  of  values  of  L.  The 

sensitivity  of  the  system  to  variations  in  L is  exhibited  by  the  data 

presented  in  Figures  2 and  3.  The  lines  going  through  the  origin  depict 

the  variation  of  m as  a function  of  m at  different  weight  fractions 
S G 

of  fuel  (Y  ) for  two  values  of  L.  In  applying  those  data  to  the  expert- 

1X5  4 

mental  results  obtained  for  the  AP/propylene  system  one  deduces  a value 

of  L = 87  ± 1 cal/gm  AP  (Table  1).  It  is  to  be  noted  that  this  value 

is  considerably  less  than  the  heat  of  sublimation  of  AP  (450  cal/gm  AP), 

an  indication  of  the  contribution  of  exothermic  reactions  in  the  condensed 

phase.  Similar  conclusions  were  drawn  in  previous  studies  of  the  tempera- 

11 

ture  distribution  in  solid  AP  during  steady-state  deflagration  and  in 

12 

a modelling  analysis  of  AP  combustion. 


The  kinetic-controlled  regime  of  the  HOFD  system  is  demonstrated  by 
the  attainment  of  the  solid  regression  rate  limit  at  high  mass  fluxes  of 


5 


rpsz 33Ctti55p&3 


gaseous  fuel.  This  condition  is  represented  by  the  intersection  of  the 


straight  lines  with  the  curves  in  the  (m  , m ) plane  of  Figures  2 and  3 

G S 

computed  for  different  values  of  L and  the  activation  energy  E for 

a single-step  reaction  of  first  order  in  fuel  and  oxidizer.  These  data 

indicate  that  with  increasing  E the  transition  to  kinetic  control  occurs 

* 

at  progressively  lower  fuel  mass  fluxes  (critical  value  = m \ However, 

G 

an  increase  in  the  initial  fuel  concentration  (Y  ) exhibits  a somewhat 

FG 

more  complex  pattern.  For  example  the  data  in  Figure  1 indicate  that 

$ 

for  E = 25  an  Increase  in  Y from  0.25  to  1.00  causes  m to  decrease 

* FG  G 

progressively  while  m appears  to  go  through  a maximum  as  observed 

4 S 

experimentally.  This  behavior  may  be  due  to  the  progressive  departure 

from  stoichiometric  conditions  (on  the  fuel-rich  side)  as  Y increases 

FG 

so  that  the  flame  cools  and  the  regression  rate  decreases. 

While  the  computations  for  Figures  2 and  3 consider  all  the  oxygen 

in  the  AP  to  be  available  for  C H -oxidation  in  the  flame  (Y  = 0.547), 

3 6 SX 

those  of  Figure  4 consider  some  of  the  oxygen  consumed  for  ammonia 
oxidation  (Y  = 0.400).  These  computations  point  to  the  marked  effect 

bA 

of  the  oxidizer  weight  fraction  at  the  solid  surface  on  the  combustion 

process.  One  also  notes  in  Figure  4 that  the  reduction  in  the  supply 

of  oxidizer  has  moved  the  maximum  regression  rate  toward  lower  values 

of  Y , as  expected  on  the  basis  of  a departure  from  stoichiometry. 

FG 

The  theoretical  model  is  applied  to  the  experimental  data*  for  the 

AP/propylene  system.  As  can  be  seen  the  two-regime  model  represents 

a satisfactory  approximation  to  the  experimental  observations  (Figure  5). 

* 

The  limiting  mass  flux  m for  each  weight  fraction  of  fuel  has  been 

S 

recorded  in  Table  1.  This  data  is  used  to  compute  the  activation  energy 
E (Appendix  A. 10).  The  value  for  E so  obtained  should  be  independent 
of  the  fuel  weight  fraction,  as  is  found  to  be  the  case  (Table  1).  The 


experimental  results  yield  an  activation  energy  of  37  x 1 kcalAiiole  for 


6 


i 


i 


the  gas-phase  reaction  associated  with  the  combustion  process  of  AP 
* 

and  propylene.  Thus  the  rate  constant  for  this  reaction  may  be  written 

13  -1  -1 

as  k = 10  exp(-37000/RT)  ccmol  sec 

IV  Conclusions 

The  opposed  flow  diffusion  flame  is  a convenient  and  precise  tool  for 

the  investigation  of  kinetic  parameters  at  high  temperatures.  In  the 

homogeneous  case  of  opposed  gaseous  jets,  the  flame  is  abruptly  extinguished 

on  the  axis  of  symmetry  when  the  reactant  flow  reaches  a critical  value; 

extinction  occurs  when  the  reactants  are  no  longer  completely  consumed  and 

therefore  act  as  diluents  carrying  heat  away  from  the  flame.  A model  of 

3 

the  flow  and  reaction  has  been  used  to  derive  overall  reaction  parameters 
from  the  opposed  flow  data. 

In  the  heterogeneous  case  one  stream  is  produced  by  gasification  of 
a condensed  phase  reactant,  the  other  originates  as  a gas.  A model  for 
the  HOFD  at  very  low  flow  rates  has  been  used  to  relate  the  ratio  of 
solid  and  gas  mass  fluxes  to  the  heat  of  gasification.  Its  constancy  as 
the  gas  composition  is  varied  indicates  that  a useful  explanation  of 
the  relation  between  regression  rate  and  heat  of  gasification  has  been 
found . 

Our  model  further  suggests  that  the  levelling-off  in  observed  regres- 
sion rate  curve  as  the  gaseous  reactant  flow  is  increased  is  due  to  in- 
complete combustion.  These  critical  conditions  can  be  related  to  the 
gas  phase  reaction  rate.  The  model  predicts  activation  energies  that 
are  independent  of  the  gas  stream  composition.  Thus  the  HOFD  offers  a 
unique  approach  to  the  evaluation  of  kinetic  parameters  of  concern  to 
complex  combustion  systems  such  as  those  encountered  in  solid  propellant 
deflagration. 

* 

For  Y = 0.400  or.e  finds  a value  of  E = 24.5  - 2 kcal/molc  AP. 

XS 


7 


Appendix  A 


EQUATIONS  OF  THE  MODEL 

A.l  Mass  and  Energy  Conservation  Equations 

For  the  one-dimensional  model  employed  in  our  analysis  we  use  the 

13 

Shvab-Zeldovitch  equations  which  govern  mass  and  energy  conservation 

in  convective,  diffusive,  reactive,  steady-state  flow.  To  reduce  the 

degree  of  complexity  of  the  problem  without  materially  affecting  the 

validity  of  the  model  the  simplifying  assumptions  are  made  that  the 

binary  diffusion  coefficient  D is  the  same  for  ail  species  and  the  Lewis 

number  is  unity,  so  that  heat  conductivity  X.  is  related  to  the  gas  density 

p and  average  specific  heat  C by  - pD  C , We  may  write  therefore: 

P P 

7.  (ov  Y - pD  7 Y ) = w 

i i l 

V.  (pv  C T - pD  C 7 T)  = -F.  h°  w 

P P i i ' 

where  v is  the  velocity  vector,  w ^ is  the  rate  of  formation  of  species  i 

in  the  single  step  reaction,  Y^  is  the  mass  fraction  and  h^0  the  heat  of 

formation  of  the  species,  and  T is  the  temperature.  If  C is  the 

pi 

specific  heat  at  constant  pressure  of  species  i then  C is  defined  by 

P 

C T = F Y |’T  C dT  , 
p i i Jo  pi 


The  stoichiometry  of  the  reaction  may  be  expressed  as : 

F v/  m - E v"  m , 

1 i i i 

where  lvi^  is  the  chemical  symbol  for  species  i and  the  and  are 
stoichiometric  ratios.  Then 


w = W (v  - v ) U)  , 
i i i i 


8 


where  is  the  molecular  weight  of  species  i and  molar  reaction  rate 
is  independent  of  i.  The  mole  fraction  of  species  i is  : 

= (Y,/W  ')/Z  (Y  /W,  ) . 

i i i k k 


Define  the  heat  of  reaction  Q by 


Q = -Z  h°  W (v"  - v7  )/w  , 

i i i i o 


where  W is  a constant  molecular  weight.  Then  the  conservation  equations 
o 

read 


L («  ) = L (a?  ) = a)  W , 

i To 


where 


or  = Y W /W  (v"  - v7  ) 

i i o i i i 


« = C T/Q  , 

T p 


and  differential  operator  L is  defined  by 
L(o)  = pv.V  a - v.p Etftf  • 

The  equation  for  conservation  of  total  mass  , 
V.(p?)  = 0 , 

has  been  used  to  simplify  L . 

A.  2 Axis  Approx imatio.i 


On  the  axis  of  cylindrical  symmetry  vector  v has  just  the  one 
component  v in  the  z-direction  so  that 


da  _ 

L(a)  = pv  t—  - T'PD?01'  . 
oz 


9 


Neglecting  radial  diffusion  gives  the  one-dimensional  equation 


L(o0 


d da; 

pD 

dz  dz 


The  equation  has  a simpler  form  if  the  axial  coordinate  is  changed 
from  z to  the  dimensionless  coordinate  y : 


y = m JZ  (1/pD)  dz  , 
o o 

where  m is  a constant  mass  flux.  Then 

da 

dy 


For  definiteness  m is  taken  to  be  m , 
o G 


the  mass  flux  at  the  gas  inlet, 


m_  - ~pv  at  z = ® . 

G 

and  the  origin  of  coordinates,  z=0,  is  put  at  the  surface  of  the  solid. 


A. 3 Flow  Field 

If  s stands  for  any  of  the  differences  (a  - a ) or  (a  - a ) then 

IT  i j 


L(s)  = 0 


or 


ms  - s = 0,  m = pv/m 


where  the  prime  denotes  differentiation  with  respect  to  y . This  can  be 
solved  for  a non-constant  solution  function  s if  m is  given  as  a function 
of  y . Conversely  if  s is  given,  m is  determined. 

In  order  that  s may  replace  y as  a coordinate,  one  requires  s/  ^ 0. 


Also,  for  convenience, 


s = 0 at  y = 0 , 
s = 1 at  y = 00  , 

so  that  s'  is  positive.  Conditions  on  m are 

m = m /m  at  y = 0 
S G 

m = -1  at  y = 10 

where  m is  the  mass  flux  at  the  surface  of  the  solid.  The  y coordinate 
S 

of  the  stagnation  point,  y , is  found  by  solving 

o 

m = 0 at  y = y . 

o 

In  the  experiment,  there  is  a cylindrical  nozzle  for  the  gas  inlet 
and  the  oxidizer  has  the  shape  of  a solid  cylinder.  The  ratios  of 
solid  diameter  to  nozzle  diameter  and  distance  to  the  nozzle  sre 
parameters  controlling  the  flow  configuration,  in  general.  However, 
both  ratios  were  taken  sufficiently  large  in  the  experiments  so  that 
no  effect  of  further  variations  was  observed.  Thus  both  ratios  are 
essentially  infinite,  and  the  geometrical  configuration  does  not  enter 
the  calculations  explicitly. 

In  order  that  m = -1  at  y =*,  function  s'  should  be  proportional 
to  exp(-y)  when  y is  large.  A function  of  this  sort  is 

s = 1 - exp(-y  - f) 


with  f bounded  at  infinity.  A convenient  choice  is  to  make  f a 
polynomial  in  [y/(y  + b)]  where  b is  a constant  taken  to  be  positive 
so  that  the  flow  region  is  free  of  singularities.  One  finds 


s'  = (i  - s>  a + f'> 


m = f /(I  + f ) - (i  + f ) 


m'  = f"'/(l  + f')  - [f"/(l  + f’’)]2  - f" 


Conditions  at  infinity  have  been  satisfied.  We  take  f to  be  quadratic: 


f = a + a x + a x , 
o 1 2 


x = y/(y  + b)  . 


Differentiation  gives 


f'  = 1(1  - x)2/b]  (a^  + ^»2x) 


f = [2  (1  - x)  /b 


I (a  - a - 3a  x) 
2 1 2 


f"'  = [6  (1  - x)Vb3]  (&i  - 2a2  + 4a2x) 


The  conditions  at  y = 0 give 


a = 0 , 

o 


a = a + (a  + b)(a  + Mb)/2  , M = 1 + m /m  , 
2 111  S G 


s = 1 + a /b  , 
o 1 


m'  = 2 [3 (b  + ai)(ai  - 2aa>  - 2 (u?  - a,)' 


2 2 2 
(b  + ai>  (a2  - a^)]/b  (b  + aj)  ( 


where  s is  the  value  of  s at  s = 0 and  m is  the  value  of  in  there, 
o S 


The  requirements  that  b and  s^'  be  positive  restrict  the  choice  of 


a and  a to  the  region  -b  < a *"  a . In  addition,  a physically 
12  12 

reasonable  flow  has  m negative  throughout.  Ensuring  this  in  general 


seems  difficult  ; however,  if  one  lakes  a2  - ~(l/2)a^  and  restricts 


9 


12 


imwijwini'i'mmgff—uM, 


a^  to  -b  < a < 0 , then  the  formulas  simplify  and  one  can  readily  see 
that  this  condition  is  satisfied. 

The  simplified  formulas  obtained  when  a = -(l/2)a  read 

2 X 


a x (1  - x/2)  , 
1 


(ai/b)(l  - x)  , 

2 4 

-3 (a^/b  )(1  - x)  , 


, 3 5 

12 (a^/b  )(1  - x) 


A. 4 Boundary  Conditions 

If  T is  the  surface  temperature  of  the  solid  and  T the  ambient 
S A 

temperature,  the  heat  flux  into  an  inert  solid  in  the  steady  state  is 


dT 

— = rn  C <T  - T)  , 
S dz  S S S A 


where  X is  the  conductivity  and  C is  the  specific  heat  of  the  solid, 
S S 

Equations  for  the  conservation  of  heat  and  species  across  the  surface 
read 


dT 


dT 


\ — = m L + A.  — I 
dz  S S dz  S 


dY. 


-PD 


dz 


ms  <YSi  ~ V * 


where  L is  the  net  heat  of  gasification  of  the  solid  and  Y is  the 

ol 

mass  fraction  of  species  i in  the  vapor  given  off  by  the  solid. 


Define  0/  to  conform  to  Q < ; 

Si  i 


a = Y W /W  (V  - v ) . 
Ci  Si  o i i i 


13 


Then  the  species  boundary  condition  becomes 


m or 

i 


”s  <ai  - V “ y 0 • 


The  heat  condition  for  the  inert  solid  reads 


°T  ’ <VV  ta/«>  * <CS  V ‘“tS  - V’  “ J = ° • 

where  the  variation  of  C with  temperature  and  concentration  ratios 

P 

has  been  neglected.  If  there  is  heat  release  by  reaction  in  the  solid 
phase, the  heat  condition  has  a different  form.  Only  inert  solid  is 
treated  here. 

At  the  gas  inlet,  the  temperature  and  species  concentrations  have 
known  values : 

ot  = a , or  = a , at  s = 1 , 

T TG  i iG 


A. 5 Species  Distributions 

The  distribution  of  species  i is  related  to  that  for  temperature 


by 


a = a - A - B s , 
i T i i 


where  A^  and  are  constants  determined  by  the  boundary  conditions. 
One  finds 


A = of  -Of 

i TS  iS 


A + B = Of  -a 
i i TG  iG 

s ' B = (m  /m  ) UL/Q)  + (C  /C  ) (or  -or  ) - -or  )]  . 
o i S G S p TS  TA  iS  Si 


These  conditions  are  used  to  determine  A , B , and  or  for  given  s . M, 

i 1 iS  o 

and  or 

TS 


14 


A. 6 Reaction  Rate  Equation 


The  Arrhenius  gas  phase  reaction  rate  expression  may  be  written  as 
n -E/R  T _ , n . 

• u)  = BT  e o FI  (X  p/R  T)  3 , 

3 o 

where  E is  the  activation  energy,  R the  gas  constant,  B and  n are 

o 

constants,  n.  is  the  order  of  the  reaction  with  respect  to  reactant  j, 

0 

and  index  j ranges  over  the  species  consumed  in  the  reaction.  Here 
j = F or  X . 


Let  W be  the  average  molecular  weight , 


W = 1/Zfl^/W  > . 

Then  X = Y W/W  so  that  the  X can  be  eliminated.  Neglecting  the 
3 3 _ 3 3 

variation  in  W allows  one  to  set  W = W.  Since  Y and  T are  proportional 

o 3 

to  the  variables  of  the  theory,  « and  ot  one  obtains 

j T 

» . B (p/s  «pyc)n-aj  . V/<vt  n , 

O T p j j 

E*  = E C /R  Q . 

P o 

The  equation  for  conservation  of  energy, 

L (a^)  = uj  w , 

simplifies  when  the  independent  variable  y is  replaced  by  s to  read 

2 2 

O'  **  + R = 0 , R = (w  W)  [pD/m^  (s')  ] , 

T G 

where  the  dots  indicate  differentiation  with  respect  to  s. 

The  dependence  of  pD  on  the  temperature  is  given  by 


PD  = o D (T/T  ) 
A A A 


15 


m 
m 


Then 


_ t i -2  n+d-En-i  -E  /^m  n n; 

R = B (s  ) or  Je  Tri(A+Bs-»)J 

T 3 3 T ’ 


b'  = (BW/m  2)  p Dl'd  (p/R  >&li  (Q/C  )n+d_l:nj  J]  (V  ')nj 
G AAA  o p j 


Boundary  conditions 


a = or 

T TC 


on  0/  read 
T 

at  s = 1 , 

(m  /m  ) [(L/Q)  + (C  /C  ) 
S i S p 


(a  - <*  )]  at  s = ° . 

T TA 


The  curve  in  the  (s,  & ) plane  representing  the  solution  lies 
inside  the  polygon  defined  by 

0*3*1,  J^or  , a A . + B s 
T T 3 3 


~._tc  lunction  R is  positive  inside  the  polygon  and  is  zero  on  the  sides, 
except  the  sides  s = 0 and  s = 1.  The  differential  equation  shows  that 
the  second  derivative  or,  for  brevity,  "curvature"  of  the  solution 
curve  is  equal  to  (-R).  Where  reaction  occurs,  rate  function  R is 
positive  and  the  solution  curve  is  convex  above,  i.e.,  arched. 


rs 

f: 


16 


I 


L. 


I 

§ 


A. 7 Approximate  Temperature  Distribution 

Possible  solution  curves  for  dimensionless  temperature  O'  as  a function 
of  distorted  distance  coordinate  s are  sketched  in  Figure  6.  Point  A repre- 
sents the  thin,  diffusion-controlled  flame  for  which  the  reaction  rate  is 
so  fast  relative  to  the  rates  for  diffusion  and  convection  that  both  react- 
ants are  completely  consumed  at  the  flame  surface.  For  this  case,  reaction 
rate  function  R is  infinitely  large  in  the  interior  of  the  polygon,  but  is 
zero  on  its  sides. 

For  finite,  decreasing  values  of  R the  reaction  zone  broadens, 
and  temperature  curves  are  found  like  those  labelled  1 to  4 in 
Figure  6,  Point  V,  where  R is  a maximum,  has  been  taken  in  the  model  as 
f..e  flame  location.  It  should  be  close  to  the  points  of  maximum  temper- 
ature and  maximum  reaction  rate. 

The  temperature  distribution  is  approximated  in  the  model  by  a 
polynomial  in  s that  fits  the  energy  conservation  equation  at  three 
points , the  two  eeges  of  the  flame  zone  and  a point  V at  the  "center" 
of  the  flame  where  the  reaction  rate  is  a maximum.  A polynomial  of 
fourth  degree  is  required  to  fit  the  second  order  equation  and  its  two 
boundary  conditions : 

2 3 4 

= b+bs+bs+bs+bs. 

T o 1 2 3 4 


At  Point  V,  we  have 


a = l.  b s 
TV  j V 

.1=0 


£ j<j  - 1)  b.  syJ  2 + Ry  = 0 


R + R 4-  j b s 
,sV  .ttjV  j V 


j-1 


17 


where,  in  the  last  equation,  the  subscript  commas  indicate  partial 
differentiation.  From  the  formula  for  reaction  rate  function  R one  finds 


R = (-  2m  + n B /Z  + n B /Z  ) R , 

,s  F F F X X X 


, 2 

R = [E  /a  + (n  + d - n - n )/ot  - n /Z  - n /Z  ] R , 
,<*t  T F X T F F X X 


Z = A + B s - O'  ,Z=A+Bs-Q'. 
F FF  T X X X T 


The  remaining  equations  and  boundary  conditions  needed  to  determine 

the  b.  depend  on  which  type  of  solution  curve  is  applicable.  If  both 
J 

fuel  and  oxidizer  are  completely  consumed,  the  curve  will  be  like  Type  1 
as  sketched  in  Figure  6 . In  the  limit  where  the  fuel  is  just  consumed 
before  reaching  the  solid  surface,  the  curve  is  of  Type  2.  The  equations 
that  follow  are  written  for  Type  1 and  apply  to  its  limit,  Type  2.  The 
other  types,  with  incompletely  consumed  fuel,  were  not  used  in  the 
comparisons  with  the  experimental  observations. 


If  point  (s  , of  ) is  at  the  inlet  edge  of  the  reaction  zone, 
X TX 


a ' = £ b s 0 

TX  j X 


a = A + B s 
TX  XXX 


^ . . J-l 

£ J b0  sx  ■ Bx 


£ j (j  - 1)  b S 
J A 


3-2 


= 0 


where  the  model  polynomial  passes  through  the  point  according  to  the 

first  equation,  the  point  lies  on  Y =0  according  to  the  second  equation, 

X 

in  the  third  equation  the  curve  is  given  the  slope  of  Y = 0 since  slope 
is  proportional  to  heat  flux  and  is  therefore  continuous,  and  in  the 


18 


fourth  equation  the  curve  is  given  zero  second  derivative  as  determined 
by  the  differential  equation  being  fitted.  At  the  other  edge  of  the 
reaction  zone  the  same  equations  are  valid  if  subscript  X is  replaced 
with  F. 


The  11  equations  above  may  be  solved  for  the  11  unknowns:  the 

5 b 's  and  the  pairs  of  coordinates  of  Points  V,  F,  and  X. 
j 


A. 8 The  Solid-Gas  Interface 

In  an  experiment  with  a known  reactive  system  the  material,  kinetic, 

and  configuration  parameters  would  be  known.  If  conditions  permitted  a 

steady-state  flame,  values  for  the  surface  temperature,  flame  location, 

and  regression  rate  could  oe  measured.  A complete  model  therefore 

requires  that  these  quantities  be  computable.  It  has  been  shown  above 

that  the  regression  rate  is  determined  in  the  model  if  the  surface 

/ 

temperature  and  parameter  are  given. 

Since  the  details  of  the  gasification  process  at  the  solid  surface 

are  not  iuiown  in  quantitative  detail,  the  model  has  been  completed  by 

10 

using  the  empirical  relation  between  surface  temperature  T,(°K)  and 

S 

regression  rate: 

2 

m,  = 1.8  exp  (14.9  - 16500/T  ) (gm/cm  -sec). 

b S 

Specification  of  s'  has  been  achieved  at  two  limiting  points  by  use 
o 

of  the  extra  conditions  valid  there.  The  thin-flame  limit  is  discussed 

in  the  next  section.  The  other  limit  is  where  the  reaction  zone  just 

reaches  the  surface  of  the  solid.  The  extra  condition,  s = 0,  is  an 

F 

indirect  specification  of  the  free  parameter.  This  limit  gives  a change 
of  flame  character  in  the  model  and  is  taken  to  coincide  with  the  ob- 
served changes  in  rate  of  change  of  regression  rate  with  inlet  flow  shown 
in  Figure  5. 


19 


A. 9 Low-Flow  Regime 


If  the  inlet  gas  mass  flux  m is  small,  a steady  flow  solution  has 

G 

very  low  reaction  rate  almost  everywhere  since  the  reaction  rate 

t -2 

function  R is  proportional  to  cu(m  s ) and  R and  s are  bounded.  This 

G 

condition  Ls  the  thin-flame  approximation  without  any  fuel  on  one  side 

of  the  flame  or  any  oxidizer  on  the  other.  In  particular,  at  the  surface 

of  the  solid,  the  slope  of  the  solution  curve  ot'  is  related  to  the 

TS 

other  parameters  and  s by 


/ / 

3 = Of  /B  . 

o TS  F 


The  pressure  variation  due  to  flow  acceleration  is 


3 - / 

-p  = pv  v = m (R  Q/p  DC)  m(npf  ) . 

z zGoAAp  T 


Thus  p goes  to  zero  with  m due  to  dimensional  considerations.  It 
z G 


seems  reasonable  that  the  nondimens ional  flow  pattern  also  contributes 
to  the  smoothing  out  of  the  pressure  variations.  It  is  therefore 


assumed  that  (rt*  ) also  goes  to  zero  with  m at  the  surface  of  the 
T G 


solid . This  gives 


/ t 

m O'  + on  /m  ) O'  = 0 . 
S TS  S G TS 


/ / 

The  flow  model  provides  an  expression  for  m^  in  terms  of  s and 


(m  /m  ) : 
S G 


m'  = - (s  + M-l ) [3M  - 3 + (M  + 5)  s'  - 2(s'  )]/3(l  - s'  ) 

So  o o 


o 


where  M = (m  /m  ) + 1 . One  may  then  solve  for  the  value  of  (m  /m  ) in 
S G S G 


the  low-flow  limit: 


m /m  = 3 [3  - or  (l  + 0)(l  + 20)1/3  [33  + a (1  + 3 ) (1  - 2&)1 
S G 


where 


ci  = ci  / b 

TS  F 


B = ci  -a  -a 
F TG  FG  TS 


3 = [L/Q  + (C  /C  ){a  - * )]/Br  . 

S p TS  TA  F 


and,  if  Tc  = Ta  , »TS  = UTA  • If  « « 1 the  relation  simplifies  to 


s = ci  ■ f v (9  + <1  m /m  )/3  . 

S G 


A. 10  Calculation  Method 


For  a specified  reaction  system  the  calculation  requires  knowledge 

of  the  species  present,  the  reaction  stoichiometry,  the  ambient  and  inlet 

temperatures,  the  weight  fraction  Y of  fuel  in  the  inlet  stream  and  the 

FG 

weight  fraction  Y of  oxidizer  in  the  solid.  Values  for  the  specific 
SX 

heats,  C and  C , and  the  heat  of  combustion  Q are  also  known  so  that 
s p 

the-  nondimensional  quantities  ct  a ./  are  determined.  In  the  low 

TA  TG  FG 

flow  limit  (Section  A. 9),  if  T = T , one  has  = <v  so  that  B is 

A C-  TS  TA  F 

known  and  the  relation  between  heat  of  gasification  I.  and  low-flow  mass- 


flux  ratio  m /m  is  fixed.  The  observed  value  of  this  ratio  determines  L. 
S G 


For  the  case  of  complete  combustion  of  the  gaseous  fuel  and  no 

fuel  in  the  solid,  one  has  Y = Y =0.  The  mass-flux  ratio  m /m 

FS  SF  S G 

is  assumed  to  remain  constant  up  to  the  limit  of  complete  combustion. 

10 

For  each  m , the  empirical  regression  relation  determines  the  surface 
S 

/ 

temperature  T . The  values  of  A^,  and  B^,  and  s are  then  computable 


21 


from  the  boundary  conditions  for  the  fuel  species  at  the  solid  surface 

given  in  Section  A. 5.  The  three  oxidizer  boundary  conditions  then  serve 

to  compute  A , B , and  ot  or  Y . the  mass  fraction  of  oxidizer  in  the 
X X XS  XS 

gas  at  the  solid  surface. 


For  specified  kinetic  parameters,  the  reaction  rate  function  R is 
now  a computable  function  of  s and  or  . The  conditions  on  the  temperature 
distribution  parameters  given  in  Section  A. 7 can  be  readily  reduced  to 


a pair  of  nonlinear  equations  for  s and  s , the  s coordinates  of  the 

F X 

edges  of  the  reaction  zone.  (At  s the  fuel  concentration  drops  to  zero; 

F 

while  at  s^  the  oxidizer  concentration  drops  to  zero.)  The  equations  have 

been  solved  for  s and  s by  a two- variable  version  of  Newton's  method. 

F X 


The  computations  were  carried  out  for  different  values  of  the 

regression  rate  m . As  the  limiting  solution  that  value  of  m was 
S S 

selected  for  which  the  reaction  zone  extended  to  the  solid  surface  so 

that  s =0, 

F 


i 

I 

t 

i 

| 


22 


Appendix  B 


APPLICATION  OF  THEORETICAL  ANALYSIS  TO  THE  AP-FUEL  SYSTEM 

To  demonstrate  the  suitability  of  the  theoretical  model  for  evalu- 
ation of  reaction  kinetics  of  a heterogeneous  combustion  system  we  have 
carried  out  a series  of  calculations  for  the  opposed-flow  diffusion  flame 
of  solid  AP-gaseous  propylene,  for  which  experimental  data  are  available. 
The  reaction  is  considered  to  involve  the  thermal  decomposition  of  AP 
during  the  gasification  process  with  subsequent  combustion  between  the 
oxygen  produced  and  the  propylene  added  in  accordance  with  the 
stoichiometry 

C H + 1/2  0 3 CO  + 3 H 0 

3 6 2 2 2 

In  the  current  analysis  the  contribution  of  fuel  from  the  subliming  solid 

(AP)  is  taken  to  be  small  compared  to  that  introduced  from  the  gaseous 

fuel  side  (CH ).  Consequently  we  do  not  consider  explicitly  the  forma- 
3 6 

tion  of  a premixed  flame  in  close  proximity  of  the  solid  surface  due  to 

the  reaction  of  the  decomposition  products  of  AP  (NH  and  HC10  , or  its 

3 6 

oxidizer  intermediates)  as  postulated  in  the  granular  diffusion  flame 
14 

model.  Such  exothermic  reactions  are  buried  in  the  gasification  term 

used  in  our  analysis  and  together  with  solid-pnase  exothermic  reactions 

contribute  to  a reduction  in  the  absolute  value  of  the  heat  of  sublima- 

15 

tion  of  AP  from  480  cal/g  AP  to  87  cal/g  AP.  However,  in  order  to 

examine  the  effect  of  oxygen  concentration  Y at  the  solid  surface 

bX 

on  the  reaction  kinetics  of  the  diffusion  flame  we  have  carried  out 

several  computer  calculations  at  two  levels  of  Y , one  for  Y = 0.547 

SX  SX 

corresponding  to  all  the  oxygen  in  solid  AP,  the  other  at  Y =0.4 

SX 

corresponding  to  some  oxygen  depletion  (due  to  reaction  with  ammonia). 

For  the  conditions  prevailing  at  the  gaseous  fuel  side  we  have  selected 
three  fuel  weight  fractions,  i.e.,  0.32,  0.60,  and  1.00  corresponding 


to  24,  50,  and  100  vol%  of  propylene,  as  employed  during  the  experimental 
study.  The  remaining  input  parameters  for  the  computer  calculations  are 
listed  in  Table  2. 

A comparison  of  the  present  two-regime  model  with  the  experimental 
4 

data  is  made  in  Figure  5.  The  model  is  fitted  to  the  data  at  the  two 
extremes  of  m near  zero  and  m at  its  limiting  value  in  . 


REFERENCES 


1.  Reviews  are  to  be  found  in  J.  A.  Steinz,  P.  L.  Stang,  and 

M.  Summerfield,  AIM  Publication  65-658,  and  J.  S.  Ebenezer, 

R.  B.  Cole,  and  R.  I.  McAlevy,  III,  Technical  Report  ME-RT  73004, 

Steven',  Inst.  Technology,  Hoboken,  N.  J.,  June,  1973. 

2.  A.  E.  Potter  and  J.  N.  Butler,  Amer.  Rocket  Soc.  J.,  29,  54  (1959). 

3.  C.  M.  Ablow  and  H.  Wise,  Combustion  and  Flame,  22,  23  (1974). 

r 4.  S.  J.  Wiersma  and  H.  Wise,  "Solid  Propellant  Kinetics.  IV. 

Measurement  of  Kinetic  Parameters  in  Opposed  Flow  Solid  Propellant 
Diffusion  Flames,"  Interim  Technical  Report,  Contract  N00014-70-C-0155 , 
Office  of  Naval  Research,  Power  Branch  (December  1973), 

5.  D.  B.  Spalding,  Amer.  Rocket  Soc.  J.,  31,  763  (1961). 

6.  F.  E.  Fendell,  J.  Fluid  Mech. , 21,  281  (1965). 

7.  L.  Krishnamurthy  and  F.  A.  Williams,  "A  Flame  Sheet  in  the  Stagnation- 
Point  Boundary  Layer  of  A Condensed  Fuel,"  paper  presented  to  Western 
States  Section,  The  Combustion  Institute,  Tempe,  Arizona,  April  1973. 

8.  A.  Linan,  "Asymptotic  Structure  of  Counterflow  Diffusion  Flames  for 

»»  ' 

Large  Activation  Energies,  Instituto  Nacional  de  Tecnica  Aeroespecial , 
Madrid  (1973)  unpublished. 

9.  C.  Guirao  and  F.  A.  Williams,  "Models  for  the  Sublimation  of  Ammonium 
Perchlorate",  Paper  69-22,  Western  States  Section,  The  Combustion 
Institute,  China  Lake,  April  1969, 

1C.  J.  F.  Lieberherr,  12th  Symposium  (International)  on  Combustion 
(1968),  533-541. 

11.  H.  Wise,  S.  H.  Inami,  and  L.  McCulley,  Combustion  and  Flame  11 , 

483  (1967). 

12.  C.  Guirao  and  F.  A.  Williams,  "A  Model  for  Ammonium  Perchlorate 
Deflagration  between  20  and  100  atm,"  AIAA  J.  Vol.  9,  pp.  1345-1356 
(1971). 

13.  F.  A.  Williams,  Combustion  Theory,  (Addison-Wesley , Palo  Alto,  1965). 

14.  M.  Summerfield,  G.  S.  Sutherland,  M.  J.  Webb,  H.  J.  Taback,  C.  P.  Hall, 
"Burning  Mechanism  of  Ammonium  Perchlorate  Propellants",  in  Solid 
Propellant  Rocket  Research,  M.  Summerfield,  ed.,(Acad.  Press,  N.Y.  1960). 

15.  S.  H.  Inami,  W.  A.  Rosser,  and  H.  Wise,  Combustion  and  Flame,  12, 

41  (1968). 


2b 


Table  2 


INPUT  PARAMETERS 


Quantity 

Symbol 

Value 

Molecular  weight-fuel 

WF 

42  gm/mol 

-oxidizer 

wx 

32  gm/mol 

Pressure  (1  atm) 

P 

0.0242  cal/cc 

Gas  conductivity  at 
ambient  conditions 

X 

A 

—5 

6.0  x 10  cal/cm-sec- °K 

-3 

Gas  density  at  ambient 
conditions 

PA 

1.25  x 10  gm/cc 

Lewis  number 

Ypa  da  5p 

1 

Ambient  temperature 

ta 

300  °K 

Inlet  temperature 

tg 

300  °K 

Solid  density 

PS 

1*8  gm/cc 

Temperature  dependence 

of  pD 

d 

0 

of  B 

n 

0 

Order  of  reaction 

for  fuel 

n 

T? 

1 

for  oxidizer 

r 

nx 

1 

Stoichiometric  coefficient 

for  fuel 

v' 

F 

1 

for  oxidizer 

V7 

X 

4.5 

Specific  heat  - solid 

"s 

0.25  cal/°K 

- gas 

n 

0.25 

Arrhenius  preexponential 

p 

B 

]0  —1  —1 
10AO  cc»mol  sec 

27 


SOLID 

OXIDIZER 


l 

i 

e 

| 


II  II 

T A-8378-48 

FIGURE  1 SCHEMATIC  CROSS-SECTION  OF  HOFD  FLAME 


£ 


28 


10*  (gm/cm 


\t  FIGURE  2 THEORETICAL  REGRESSION  RATE  ms  VERSUS  INLET  FLOW  mG  LINES  AND 

| COMPLETE  COMBUSTION  LIMIT  CURVES  (Yx$  = 0.547,  L = 87  cal/gmAP) 


29 


10 


2 2 

mG  *10  (gm/cm  -*»c) 


TA-8378-47 


FIGURE  3 l.'EORETICAL  REGRESSION  RATE  m$  VERSUS  INLET  FLOW  mG  LINES  AND 
COMPLETE  COMBUSTION  LIMIT  CURVES  (Yxs  = 0.547.  L = 85  cal/gmAP) 


30 


I 


0 | 1 * 

TA-8378-44 

FIGURE  6 NONDIMENSIONAL  TEMPERATURE  aT  AS  A FUNCTION 
OF  DISTORTED  DISTANCE  s 


33 


32 


0 


1 


% 


TA-S378-44 

FIGURE  6 NONDIMENSIONAl  TEMPERATURE  oT  AS  A FUNCTION 
OF  DISTORTED  DISTANCE  s 


33